Sei sulla pagina 1di 12

Available online at www.sciencedirect.

com

Journal of the European Ceramic Society 28 (2008) 1557–1568

Analysis and interpretation of refractory microstructures in


studies of corrosion mechanisms by liquid oxides
J. Poirier a,∗ , F. Qafssaoui b , J.P. Ildefonse c , M.L. Bouchetou a
a
CRMHT, 1D, avenue de la Recherche Scientifique 45071 Orléans cedex 2, France
b Moulay Ismaı̈ University, Faculty of Science, B.P. 4010 Zitoune, 50000 Meknes, Morocco
c Polytech’Orléans, 8 rue Léonard de Vinci, 45072 Orléans cedex 2, France

Received 1 August 2007; received in revised form 29 September 2007; accepted 7 October 2007
Available online 4 January 2008

Abstract
The refractories must not only resist high temperatures but also corrosion by liquid oxides. This corrosion involves phenomena of dissolution and
precipitation of new crystalline phases. The study of the microstructures of corroded refractories provides essential information. However, the
interpretation of the microscopic observations is difficult. Indeed, because of the crystallization of liquid glasses during cooling, the mineral phases
observed at room temperature are not representative of those observed at high temperature. The concept of local thermodynamic equilibrium and
the use of the phase rule makes it possible to interpret the microstructures of corroded refractories, to explain the observed mineral zonation and
to quantify the composition of the liquid phase at high temperature from chemical profiles established by SEM. Experimental data from corrosion
of high alumina refractories will illustrate and validate this theoretical approach.
© 2007 Elsevier Ltd. All rights reserved.

Keywords: Refractories; Corrosion; Microstructure; Thermodynamic equilibrium; Electron microscopy

1. Introduction • The microscopic observations and the analyses are carried


out at room temperature. They are not representative of the
The refractories are ceramics used to line many industrial mineral and vitreous phases existing at high temperature.
high temperature furnaces. These materials are subjected to com- • During cooling, new solid phases appear by crystallization
plex degradations such as thermal shock, erosion or chemical of liquid oxides. The composition of the vitreous phases also
corrosion which can occur separately or together. evolves with the temperature (see Fig. 1).
Corrosion by liquid oxides is one of the most severe modes
of degradations which limit the lifetime of the refractory linings. Consequently, the information obtained is often limited and
The examination of the microstructures1 of refractories after gives already known conclusions.
use is extremely useful to evaluate the corrosion resistance of In this paper, we will present a method to analyse and interpret
various refractories and to determine the mechanisms of chem- the microstructures of refractories after use. Typical examples
ical attack thus making it possible to propose new ways of of corrosion of high alumina-based refractories will be pre-
improvements for the formulation of refractories. sented.
However, the microstructures of corroded refractories are The experimental results, which are not well explained, will
very difficult to interpret. The reasons are as follows: be re-analysed and clarified.
By using thermodynamic data and phase diagrams, the study
• The refractories are multi-component and heterogeneous of the microstructures will enable us to deduce:
ceramics. Consequently, they have complex microstructures.
• The proportion and the composition of the high temperature
intergranular liquid phase.
∗ Corresponding author. Tel.: +33 2 38 25 55 14; fax: +33 2 38 63 81 03. • The gradients of the chemical properties between the hot face
E-mail address: Jacques.Poirier@univ-orleans.fr (J. Poirier). and the back of the refractory lining: the chemical profile of

0955-2219/$ – see front matter © 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jeurceramsoc.2007.10.012
1558 J. Poirier et al. / Journal of the European Ceramic Society 28 (2008) 1557–1568

Fig. 1. SEM micrographs of andalusite based refractory in (a) naturally cooled and (b) quenched crucibles (laboratory test: T = 1600 ◦ C, 50 wt% Al2 O3 –50 wt% CaO
slag and crucible method).

composition of glasses, the evolution of the viscosity of the


liquids.
Table 1
Formulations of andalusite and bauxite bricks (manufacturer data)
2. Experimental procedure and examination of
Raw material Size (mm) Andalusite Bauxite brick
microstructures
brick (wt%) (wt%)

Two high alumina refractory materials based on andalusite Randalusite 1–4 30


Kerphalite KA 0.3–1.6 35
and bauxite have been selected for the purpose of this research.
Kerphalite KF 0–0.16 12
Their formulations and principal characteristics are indicated Kerphalite KF 0–0.055 5
in Tables 1 and 2. The typical microstructures of the original Chinese bauxite 1–4 32
bricks are shown in Figs. 2 and 3. For the andalusite refractory, Chinese bauxite 0–1 33
a fine matrix composed of mullite needles in a silica-rich glass Chinese bauxite 0–0.16 9
Calcined alumina 14 14
binds the coarse grains of mullitized andalusite converted into
Clay RR40 4 9
a mullite/glass composite.2 The binding glass has a composi- Plastifier 3
tion similar to that resulting from the mullitization of andalusite
(≈70% SiO2 , ≈25% Al2 O3 , ≈2% Fe2 O3 , ≈1% TiO2 and ≈1% Total 100 100
K2 O).
J. Poirier et al. / Journal of the European Ceramic Society 28 (2008) 1557–1568 1559

Table 2
Chemical composition, density and apparent porosity of andalusite and bauxite bricks
Refractories Composition (wt%) Mineral phases Density (g/cm3 ) Apparent porosity (%)
Al2 O3 SiO2 CaO MgO Fe2 O3 TiO2 Na2 O K2 O

Andalusite brick 64.1 33.4 0.1 0.1 0.7 0.2 0.1 0.2 M***, A* 2.65 11.09
Bauxite brick 78.6 11.2 0.4 0.4 1.7 3.3 0.2 0.3 A***, M**, TiO2 *, (Al, Fe)2 TiO5 * 3.24 16.17

Mineral phases: A → corundum, M → mullite (***: major; **: mean; *: minor).

Fig. 2. Backscattered electron images of the typical microstructure of andalusite brick showing the mullite–glass composite and the bonding matrix made from
silica-rich glass: (a) low magnification (50×) and (b) high magnification (1000×). M: mullite, G: glass and P: pore.

For the bauxite refractory, coarse grains are composed of • the penetration zone,
corundum with some tialite (titanium aluminate containing some • the unaffected refractory zone.
iron). The fine matrix contains corundum, mullite and a glassy
phase. The intergranular glass is very fine and has a complex 2.1. The slag zone
chemical composition (≈45% SiO2 , ≈20% Al2 O3 , ≈8% Fe2 O3 ,
≈8% P2 O5 , ≈8% TiO2 , ≈5% CaO, ≈2% MgO, ≈2% K2 O, After cooling, a glassy slag layer is observed at the bottom
≈1% Na2 O). of the crucible. Its composition changes with a decrease in CaO
Corrosion tests, using the static crucible method, have been and an enrichment in SiO2 . The Al2 O3 content remains nearly
carried out at 1600 ◦ C, in an electric furnace, incorporating an constant with andalusite bricks and shows a small increase in
elevating hearth, allowing for the loading and unloading of the the case of bauxite bricks.
crucible. Crucibles of 100 mm× 100mm × 60 mm in size were
cut-out from the bricks, previously fired at 1550 ◦ C for 12 h. 2.2. The precipitation zone
After 6 h of corrosion tests, the crucibles were rapidly quenched
in water in order to avoid partial crystallization of the liquid The precipitation zone is formed by a succession of
phase during cooling. monomineral layers of crystals surrounded by glass. The glass
These laboratory tests were performed with an Al2 O3(50 wt%) – content may be high and is locally variable (30–80%).
CaO(50 wt%) slag (labelled AC) which is close to that of a The same mineral succession can be observed in bauxite and
ladle slag allowing production of Al killed carbon steels. This in andalusite bricks. From the penetrated zone to the slag zone,
synthetic slag was prepared by mixing, in appropriate propor- three successive monomineral layers are observed:
tions, powders comprising calcium carbonate (CaCO3 > 98%,
Alfa Aesar) and reactive alumina (CT300SG, Al2 O3 > • corundum Al2 O3 ,
99.8%). • calcium hexaaluminate CaAl12 O19 (CA6 ),
The post-mortem analysis of the crucibles after testing • calcium dialuminate CaAl4 O7 (CA2 ).
(Figs. 4 and 5) permitted the identification of four zones3 :
The texture and the shape of the layers clearly indicates that
• the slag zone, the crystals were precipitated slowly from a liquid4,5 (they do
• the precipitation zone, not result from a crystallization during cooling). Corundum from
1560 J. Poirier et al. / Journal of the European Ceramic Society 28 (2008) 1557–1568

Fig. 3. Backscattered electron images of the typical microstructure of the bauxite brick showing the porous nature of bauxite and a bond matrix made from a mix of
corundum (A), mullite (M), aluminum titanate (AT), titania (T) and glass (G): (a) low magnification (25×) and (b): high magnification (500×) → microstructures
of a grain of bauxite transformed into corundum, (c) high magnification (1000×) → the bonding matrix (c; 1000×). AT: (Al,Fe)2 TiO5 , T: TiO2 and P:
pore.

the first layer shows well-formed crystals that differ clearly 3. Interpretation of corroded microstructures
from those of the transformed bauxite. The CA6 phases are
still present, but in the case of an important corrosion of the 3.1. Equilibrium aspect
refractory, the CA2 phases are absent.
In the zone of corrosion, the heterogeneous system refrac-
2.3. The penetration zone tory/slag is open and exchanges matter with the surrounding.
The concept of thermodynamic equilibrium applies to a closed
The viscous slag invades the matrix by capillary penetration system which exchanges only heat and energy. It requires the
and mixes with the pre-existing intergranular liquid by diffusion. chemical potential of each component in all phases to be equal.
The penetration of the corrosive liquid phases leads to the super- When a system is open, equilibrium is not total, but it
ficial dissolution of the aggregates, so changes affect mainly the can be applied on the local scale. The concept of “local
matrix with a progressive increase in the glass phase. For the equilibrium” is well established in the literature6,7 and is
andalusite brick, the limit between the penetrated zone and the a fundamental assumption in models of diffusion-controlled
precipitation zone is more regular than for the bauxite brick. At reaction.
that zone limit, dissolution of mullitized andalusite grains seems In a solid–liquid refractory system, chemical exchanges occur
to be completed by the corrosive liquid, but large bauxite grains through the liquid phase (diffusion, infiltration and percola-
converted to corundum may remain beyond in the precipitation tion). When solid–liquid reactions (dissolution, precipitation)
zone. are faster than chemical transport, equilibrium motion may be
locally applied:
2.4. The refractory zone
• the solid phases are locally in equilibrium with the liquid
This zone is composed of unaffected bauxite or andalusite which surrounds them;
refractories • the Gibbs’ phase rule may be used;
J. Poirier et al. / Journal of the European Ceramic Society 28 (2008) 1557–1568 1561

Fig. 4. Microstructure of corroded andalusite refractory by Al2 O3 –CaO slag attack.

• the chemical mobility entails concentration gradients, which The phase rule applies only to equilibrium states of a system,
are the driving force of corrosion and will tend to bring the which require both homogeneous equilibrium within each phase
system back towards global equilibrium. and heterogeneous equilibrium between co-existing phases. The
phase rule does not depend on the nature and amounts of the
Depending on its local composition, the liquid dissolves the solid phases present, but only on their numbers; nor does it give
phases which are not in equilibrium with it and precipitates new information concerning rates of reactions. Non-conformity with
phases after it becomes saturated. the phase rule is the proof that equilibrium conditions do not
exist.
3.2. Use of the phase rule Data from microstructural examinations, described above,
have showed that a succession of zones is observed: the initial
The degree of freedom F (variance) of the system is given by refractory, the penetration zone, the precipitation zone and the
the phase rule of Willard Gibbs.8 The following equation gives remnant slag. Furthermore, the precipitation zone may contain
the usual mathematical form of the phase rule: one or several mono mineral layers.
Consider as an example the corrosion of a 3:2 mullitized
F =C+2−P (1) andalusite-based refractory by an Al2 O3(50 wt%) –CaO(50 wt%)
model slag, at 1600 ◦ C.
C: number of components of the system;P: number of phases
present at equilibrium;2: number of environmental factors (tem-
perature and pressure). 3.2.1. The initial refractory
When the temperature and the pressure are fixed, the Gibbs’ The initial refractory is composed of a solid phase
phase rule reduces to the following equation: (3Al2 O3 –2SiO2 ) and a liquid phase (a silica-rich glass) which is
saturated with mullite (P = 2). It contains only Al2 O3 and SiO2
F =C−P (2) (C = 2). The degree of freedom F (variance) is equal to 0.
1562 J. Poirier et al. / Journal of the European Ceramic Society 28 (2008) 1557–1568

Fig. 5. Microstructure of corroded bauxite refractory by Al2 O3 –CaO slag attack.

Consequently, the composition of the liquid phase is constant can exist only at a given distance from the initial interface. Lastly,
and is given by the saturation solubility of the mullite in the liquid when only one solid phase is present, the variance is equal to 1,
phase. and the composition of the liquid can vary.
The new phases which crystalline during dissolution–
3.2.2. The penetration zone precipitation processes9 consist of different successive
The refractory is partially permeated by the slag. When the monomineral layers separated by sharp boundaries.
lime diffuses through the interstitial liquid, the number of com-
ponents C increases to 3 whereas the number of phases P remains 3.3. Prediction of reaction products at refractory/slag
equal to 2. The degree of freedom becomes equal to 1, and interface
the composition of the liquid is not constant any more. The
concentrations of alumina, lime and silica vary in the liquid. At the initial stage of the reaction, on both sides of the initial
interface, there is liquid slag (50% CaO and 50% Al2 O3 ) and a
3.2.3. The precipitation zone siliceous liquid which impregnate the mullite aggregates of the
In this zone, located between the slag and the zone of impreg- andalusite-based refractory. This initial situation is described
nation, newly crystalline phases precipitate during the reactions in Fig. 6. For more simplicity, consider the example of a slag
refractory/slag. The liquid contains three oxides Al2 O3 , CaO, containing the element A and a refractory containing the element
SiO2 ; the number of components C is equal to 3. The degree of B. The slag and the refractory are coupled according to a plane
freedom F can never be less than 0 under invariant conditions. initial interface.
Consequently, the number of phases P cannot exceed 3: two In the slag, all are liquid, the concentration of element A is
solids and one liquid. When two solid phases coexist, the vari- A ), and that of the element B is equal to 0. In the refrac-
high (CSlag
ance is equal to 0 and the composition of the liquid is constant. tory, composed of the phase B, there is a little liquid with a high
However, because of diffusion and of the chemical gradients, concentration of B (CBB = solubility of B in the liquid), while the
the concentration of an element in the liquid varies regularly concentration of A is equal to 0.
depending on the distance and the time period. At the end of a On both sides of the initial interface, there is a “step” of
time t, the conditions of equilibrium between two solid phases chemical concentration. This situation is unstable, and the step
J. Poirier et al. / Journal of the European Ceramic Society 28 (2008) 1557–1568 1563

Fig. 6. Initial stage of corrosion between a refractory and a liquid slag.

will be transformed into a gradient of concentration: element A Fig. 8. Evolution of the corrosion of a solid refractory by liquid slag. Disappear-
of the slag will migrate towards the liquid of the refractory, and ance of the monomineral zone of AB when the concentration of A is decreasing
element B of the refractory will migrate towards the slag. Fig. 7 in the slag.
shows the evolution of the reaction.
Consider that, in the system A–B, exist the different phases Thus, there is the formation of a monomineral zone of AB2 .
A, AB, AB2 , and B. The concentrations at equilibrium, in the Beyond this limit, the phase AB precipitates and forms a sec-
liquid, are as follows: ond monomineral zone. The latter is gradually dissolved in the
slag.
• solubility of the phase B in the liquid slag: CBB Various monomineral zones separated by boundaries are
• concentrations, respectively of A and B in the liquid phase in obtained which move gradually while advancing on refractory
equilibrium with the assembly B + AB2 : CAB A
2 /B
and CAB
B
2 /B
material.
• concentrations, respectively of A and B in the liquid phase Theoretically, in a semi-infinite system, the number of
in equilibrium with the assembly AB2 + AB: CAB/AB A
2
and monomineral zones which are formed is equal to the number of
CAB/AB
B
2
. intermediate compounds likely to be formed between the phases
A and B. In practice, in the laboratory tests, the quantity of slag is
As long as the concentration of A in the liquid remains lower limited and its composition changes. The migration of calcium
than CABA
2 /B
(in equilibrium with the phases A + B), it will be towards refractory involves a decrease in the concentration of
localized in the zone of infiltration. The increase in the concen- this element in the slag which can be lower than the concentra-
tration of A in the liquid and the decrease in B which diffuses tion CAB/AB
B
2
in equilibrium with the assembly AB + AB2 . Phase
towards the slag, involves a dissolution of phase B. AB is not stable any more, and the corresponding monominerale
When the concentration in A exceeds CAB A
2 /B
, the phase AB2 zone disappears as Fig. 8 shows.
becomes stable and precipitates while phase B disappears by The Al2 O3 –CaO–SiO2 phase diagram9 makes it possible to
dissolution. Phase AB2 remains the only stable solid phase as determine which are the phases likely to appear during the cor-
long as the concentration of A in the liquid remains lower than rosion of high alumina by Al2 O3 –CaO. It also makes it possible
CAB/AB
B
2
. to predict the succession of the reactional zonations for each
kind of refractories:

• mullitized andalusite-based refractory: slag/dialuminate of


calcium/hexaaluminate of calcium/corundum/mullite
• bauxite-based refractory: slag/dialuminate of calcium/
hexaaluminate of calcium/corundum
• alumina-based refractory: slag/dialuminate of calcium/
hexaaluminate of calcium/corundum

For example, Figs. 9 and 10 show microstructures of the pre-


cipitation zones of bauxite and andalusite refractories after
corrosion. The alteration of the refractory microstructures can be
explained by dissolution–precipitation processes inside a liquid
phase.5 Several mineral layers can be observed.
The same succession will be observed for all the types of slag
as long as their initial composition does not differ too much from
Fig. 7. Evolution of the corrosion of a solid refractory by liquid slag. Formation
of a zone of impregnation, reactional monomineral zones and displacement of a ratio C/A + S close to 1. This zonation corresponds to the max-
the interfaces. imum number of observable monomineral zones, knowing that
1564 J. Poirier et al. / Journal of the European Ceramic Society 28 (2008) 1557–1568

Fig. 9. Corrosion of bauxite brick by Al2 O3 –CaO slag; backscattered electrons SEM micrographs. (A) Transition between the three monomineral layers of the
precipitation zone and (B) neogenic layer made of large well-formed corundum crystals developed at the expense of small badly formed corundum issued from the
transformation of a bauxite grain. C: corundum; CA2 : calcium dialuminate; CA6 : calcium hexaaluminate. The light-grey background represents the high-temperature
liquid phase partially recrystallized during cooling.

Fig. 10. Corrosion of andalusite brick by Al2 O3 –CaO slag; backscattered electrons SEM micrographs. (A) Transition between the calcium hexaaluminate (CA6 )
and the corundum layers and (B) transition between the corundum layer of the precipitation zone and the refractory (mullite + glass) C: corundum; CA2 : calcium
dialuminate; CA6 : calcium hexaaluminate; M: mullite. The light-grey background with the needle-like crystals localized in the corundum layer in (A) represents the
high-temperature liquid phase partially recrystallized during cooling and converted to anorthite (CAS2 ) and residual glass.

the evolution of the composition of the slag during the reaction • In addition, the existence of monomineral zones with newly
can make the most external zones to disappear. formed solid phases in the liquid make the analyses by map-
ping non-representative, since, they integrate not only the
4. Determination of the composition profiles of the crystals formed during cooling and residual glass, but also
liquid phase at high temperature the crystals newly formed.

4.1. Method One of the possibilities consists in carrying out a fast quench at
the end of the corrosion test to limit the crystallization of the
The gradients of concentration, which exist in the liquid phase liquid during cooling. Under these conditions, the only crys-
at high temperature are the driving force behind corrosion pro- tallized phases are due to corrosion and the microanalyses of
cesses at high temperature. Unfortunately, two phenomena make glass between the crystals are representative of glass at high
it difficult to determine the composition of this liquid. temperature.
Concurrently with this direct method, an indirect method can
• On the one hand, the partial crystallization of the liquid during be implemented. It consists in using the analyses (obtained by
cooling does not make it possible to assimilate the compo- mapping) which are carried out on the totality of the corroded
sition of residual glass (which can be determined by EDS zone. The results are interpreted using the phases diagrams and
analysis) with that of the liquid which existed at high tem- the concept of “local equilibrium”.
perature. The only representative analysis would be an EDS Fig. 11 shows the principle of the method. Within the frame-
mapping of a zone including residual glass and the crystals work of the “local equilibrium concept”, the information which
formed during cooling. is provided by the phase diagrams is limited:
J. Poirier et al. / Journal of the European Ceramic Society 28 (2008) 1557–1568 1565

Fig. 11. Determination of the composition of the liquid at 1600 ◦ C for an analysed zone represented by point M (thick line: 1600 ◦ C isotherm line).

• the point L is representative of the composition of the liquid 4.2. Validation of the method and results
at 1600 ◦ C;
• the composition represented by the point M corresponds to Composition profiles (chemistry as a function of dis-
an assemblage of a liquid and a solid phase at 1600 ◦ C. tance from the initial interface), etablished for the mullitized
andalusite based refractory in contact with a Al2 O3 –CaO slag,
by the direct method (quench of the liquid phase at the end of the
The diagram Al2 O3 –CaO–SiO2 gives no direct information corrosion test) and the indirect method (concept of local equi-
on the composition, neither of the liquid, nor of the solid cor- librium applied to phase diagram) are given in Fig. 12. The data
responding to the point M. It cannot be used like a diagram obtained by these two methods are in agreement.
of thermodynamic equilibrium. However, its use as a ternary Across the modified area (zone II and III), the liquid com-
diagram of composition remains valid. A surface analysed by position presents gradients of concentration connecting the
mapping with the scanning electron microscopy (SEM) consists composition of the interstitial liquid in the refractory to the slag.
of liquid and newly formed solids during corrosion at 1600 ◦ C. The chemical exchanges occur through the liquid phase.
Those can be distinguished from the phases formed by the crys- Depending on its local composition, the liquid dissolves the solid
tallization of glass during cooling. phases which are not in equilibrium with it and precipitates new
For an analysed zone, represented by the point M, in which phases after it becomes saturated.10
the solid phase is the corundum (point S), the liquid at 1600 ◦ C Fig. 13 shows an other compositional profile of the liq-
is represented by the point L obtained by the intersection of uid phase associated with the zonation of mineral solid phases
the segment SM with the isotherm 1600 ◦ C of the diagram. The present at a temperature of 1600 ◦ C (mullitized andalusite refrac-
proportion of liquid is given by the ratio MS/LS, and the com- tory, Al2 O3 –CaO slag and quench).
position of this liquid is deduced from the position of the point For each zone, the equilibrium constants, the number of
L in the triangle of composition. phases in equilibrium and the variance of the system are deter-
In a certain number of cases, the analysed surface is situ- mined.
ated on several zones of reaction and contains several phases:
for example, mullite + corundum, or corundum + CaO. In strictly
speaking, it would be necessary to consider the relative propor- 4.2.1. Zone of the initial refractory
tion of the two phases and to place the point representative of the At this depth, CaO does not infiltrate. The liquid is com-
solid in proportion on the segment which joins these two solids. posed of Al2 O3 and SiO2 . As [Al2 O3 ] + [SiO2 ] = 1, alumina and
1566 J. Poirier et al. / Journal of the European Ceramic Society 28 (2008) 1557–1568

Fig. 12. Compositional profile of the intergranular-liquid phase in the andalusite refractory (a: experimental values, b: obtained with the Al2 O3 –CaO–SiO2 phase
diagram).

silica contents are fixed. The liquid is saturated with mullite, gradually dissolves until it totally disappears at the boundary
the saturation is defined by the equilibrium constant KM of the separating the zones of penetration and precipitation. The com-
reaction: position of the liquid phase is not constant any more and the
degree of freedom becomes equal to 1.
Al6 Si2 O13  = 3[Al2 O3 ] + 2[SiO2 ]
KM = [Al2 O3 ]3 ·[SiO2 ]2 (I) 4.2.3. Boundary between the impregnation zone and the
precipitation zone
· means “in the solid state” and [·] “dissolved in a liquid phase” Two mineral phases coexist: residual mullite and neoformed
The liquid composition is then constant and the degree of corundum. The liquid phase is simultaneously in equilibrium
freedom is 0. with corundum and mullite (C/M equilibrium). The corundum
saturation is expressed by the reaction constant:
4.2.2. Impregnation zone
CaO infiltrates the interstitial liquid. As a consequence, the Al2 O3  = [Al2 O3 ] KC = [Al2 O3 ] (II)
liquid composition will not be in equilibrium with mullite, which Al2 O3 and SiO2 contents are both constant:
 1/2
−1 KM
[Al2 O3 ]C/M = KC and [SiO2 ]C/M = KC
KC
As the liquid contains only three species, the concentration
of CaO can be determined by means of the following equation:

[Cao]C/M = 1 − ([Al2 O3 ]C/M + [SiO2 ]C/M )


It thus appears clearly that the contents of the three con-
stituents of the liquid are fixed at the boundary between the
zones of precipitation and penetration.
The degree of freedom is equal to 0.
This result is consistent with data obtained from composition
profiles previously established (either by direct analysis of glass
or using a CaO–Al2 O3 –SiO2 diagram) which revealed the exis-
tence of a constant composition at this interface whatever the
nature of the slag. In the precipitation zone, a first monomineral
layer containing corundum formed during corrosion at 1600 ◦ C
is observed. Equilibrium of liquid with corundum is defined by
reaction (II) mentioned above. It follows that Al2 O3 content is
fixed in this zone ([Al2 O3 ]C = KC ), and concentrations of CaO
Fig. 13. Another compositional profile of the liquid phase associated with the and SiO2 must inversely vary because their sum has to remain
zonation of the mineral solid phases present at 1600 ◦ C (mullitized andalusite constant. It must be noted that corundum is really stable only
refractory, Al2 O3 –CaO slag, quench). near the surface of the refractory where it precipitates, while
J. Poirier et al. / Journal of the European Ceramic Society 28 (2008) 1557–1568 1567

Fig. 14. Viscosity profiles of the liquid phase in the tested andalusite (a) and bauxite (b) refractories, according to Urbain’s model. (I) slag zone, (II) precipitation
zone, (III) penetration zone, and (IV) unaffected refractory. Initial slag–refractory interface is located at d = 0 mm. Note that zone IV in (b) is less wide than in (a)
because of deep penetration of slag in brick B.

far from the corrosion front, it dissolves to maintain a constant Viscosity η was calculated according to Urbain’s model11 by
Al2 O3 content in the liquid and thus to make up for the increase the following equation: ATexp(103 B/T), A and B are two param-
in the CaO content. eters depending on the composition of the glass. The viscosity
of the intergranular liquid phase is shown in Fig. 14 for (a) and
4.2.4. Precipitation zone (b) andalusite and bauxite refractories.
EDS analysis of the glass in the two refractories integrates
• Monomineral layer of corundum
all the elements of the slag–refractory system including the
Along the CA6 layer, concentrations of Al2 O3 , CaO and
impurities. For bauxite refractory, in which there is no resid-
SiO2 are variable. The degree of freedom is equal to 1.
ual slag zone (absence of zone I corresponding to the part
• Boundary between the corundum and the hexaaluminate of
of refractory replaced by slag), a low viscosity of the liquid
calcium layers
phase, due to the presence in a high content of minor elements
At the interface between layers of corundum and CA6 , the
(amounting∼20–30%) such as Fe2 O3 , P2 O5 , TiO2 , alkaline and
liquid is saturated with these two phases, thus corresponding
earth–alkaline oxides, is observed along the modified part of the
to equilibrium between reaction (II) and (III):
refractory (area extending from the unaffected zone to the slag
< CaAl12 O19 > = 6[Al2 O3 ] + [CaO] KH zone). This fact explains why the penetration zone is wider in
bauxite than in andalusite refractory. Furthermore, the fraction
= [Al2 O3 ]6 ·[CaO] (III) of impurities coming from raw materials may be negligible in
The concentrations of CaO and Al2 O3 are defined as fol- the andalusite brick, so that the liquid phase is considered to be
lows at this interface: predominantly CaO + Al2 O3 + SiO2 .
For andalusite refractory, the liquid viscosity is constant
KH along the slag zone (zone I;μ ≈ 1 Pa s) as well as in the unaf-
[Al2 O3 ]C/H = KC and [CaO]C/H = fected refractory (zone IV; μ ≈ 1570 Pa s). From the unaffected
KC6
zone, a considerable decrease in the liquid viscosity is observed
As [Al2 O3 ] + [CaO] + [SiO2 ] = 1, the SiO2 concentration up to the precipitation zone. The liquid becomes more fluid near
is also determined at the interface. The degree of freedom is the corrosion front and easily infiltrates the grain boundaries
equal to 0. leading to the dissolution of most present fine phases. As a result,
• Monomineral layer of hexaaluminate of calcium the dissolved phases become part of the local liquid, thus increas-
Along the CA6 layer, concentrations of Al2 O3 , CaO and ing the liquid content. However, an explanation for the lack of
SiO2 are variable. The degree of freedom becomes equal to deep slag penetration in andalusite refractory lies in the fact that
1. a high amount of viscous liquid (silica-rich with a lower content
• Boundary between the precipitation zone and the slag zone of secondary oxides) is formed in the part of the refractory in
CA6 disappears at the boundary interface. Beyond this contact with the slag, which slowed down the penetration of the
interface, there is no more liquid phase. slag.

4.3. Viscosity of the liquid phases 5. Conclusion

The gradient of the chemical composition of the intergranular The corrosion mechanisms of refractories by liquid oxides are
liquid phase plays a fundamental role. The viscosity of the liquid far from being completely known. The study of microstructures,
phase is the second parameter, which influences the migration which improves our knowledge of corrosion, is extremely use-
of the species in this liquid. ful to determine the mechanisms of chemical attack. However,
1568 J. Poirier et al. / Journal of the European Ceramic Society 28 (2008) 1557–1568

the microstructures of corroded refractories are very difficult to References


interpret.
In this paper, an approach based on the concept of local ther- 1. Lee, W. E., Characterisation of corrosion mechanisms in refractories by
modynamic equilibrium and the use of the phase rule is proposed post-mortem microstructural analysis. Br. Ceram. Proc., 1997, 57, 7–15.
2. Ildefonse, J. P., Gabis, V. and Cesbron, F., Mullitization of andalusite in
to analyse the microstructures. refractory bricks. Key Eng. Mater., 1997, 132–136, 1798–1801.
An experimental corrosion study of high-alumina refracto- 3. Qafssaoui, F., Rôle et effet de l’andalousite sur le comportement à la cor-
ries by an alumina–lime model slag is carried out to illustrate rosion des céramiques réfractaires à haute teneur en alumine par les laitiers
this approach. The post-mortem analysis of microstructures sidérurgiques, PhD Thesis. University of Orleans, France, 2004.
after testing, revealed the development of a penetration zone, 4. Zhang, S., Rezaie, H. R., Sarpoolaky, H. and Lee, W. E., Alumina dissolution
into silicate slag. J. Am. Ceram. Soc., 2000, 83(4), 897–903.
followed by a precipitation zone formed of a succession of 5. Guha, J. P., Reaction chemistry in dissolution of polycrystalline alumina in
monomineral layers and show the existence of composition gra- lime–alumina–silica slag. Br. Ceram. Trans., 1997, 96(6), 231–236.
dients in the interstitial liquid. The local chemical equilibrium 6. Thompson Jr., J. B., Local equilibrium in metasomatic processes. In
concept makes it possible to explain the observed mineral zona- Researches in Geochemistry, Vol 1, ed. P. H. Abelson. John Wiley & Sons,
tion, and to quantify the proportion and composition of the New York, 1954.
7. Mueller, R. F., Mobility of elements during metamorphism. J. Geol., 1967,
liquid at high temperature from chemical profiles, established 75, 565–582.
by analyzing successive zones by means of scanning electron 8. Gibbs, J. W., Equilibrium of heterogeneous substances. Trans. Conn. Acad.
microscopy. Sci., 1874, 3, 108–248;
The composition profiles are compared to those obtained Gibbs, J. W., Equilibrium of heterogeneous substances. Trans. Conn. Acad.
by direct analysis of the interstitial glass in refractories after Sci., 1878, 3, 343–524.
9. Gentile, A. L. and Foster, W. R., Phase Diagrams for Ceramists. The Amer-
quenching. ican Ceramic Society, Columbus, OH, USA, 1963, p. 220.
The viscosity profiles, estimated by means of Urbain’s model 10. Qafssaoui, F., Poirier, J., Ildefonse, J. P., Hubert, P. and Benyaich, F.,
from the composition profiles of the liquid phase, showed that Microstructural and physicochemical studies of corroded high-alumina
the viscosity of the interstitial liquid determines the extension refractories. Silicates Ind. Ceram. Sci. Technol., 2005, 70(7–8), 109–117.
of the penetration zone in a given refractory. 11. Urbain, G., Viscosity estimation of slags. Steel Res., 1987, 58(3), 111–
116.
These results clearly validate this thermodynamic approach 12. Poirier, J., Bouchetou, M. L., Qafssaoui, F. and Hubert, P., Corrosion of high
which offers new prospects to develop more corrosion-resistant alumina refractories by Al2 O3 /CaO slag under thermal cycling conditions.
refractories.12 Interceram. Int. (Part 1 and Part 2), 2006, 55(4), 270–272, and 348–351.

Potrebbero piacerti anche